首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The first direct equilibrium dialysis titration of the blood coagulation protein bovine prothrombin fragment 1 with Mg(II) is presented. Fragment 1 has fewer thermodynamic binding sites for Mg(II) than Ca(II), less overall binding affinity, and significantly less cooperativity. Several nonlinear curve fitting models were tested for describing the binding of fragment 1 with Mg(II), Ca(II), and mixed metal binding data. The Mg(II) data is represented by essentially five equivalent, noninteracting sites; for Ca(II), a model with three tight, cooperative sites and four "loose", equal affinity, noninteracting sites provides the best model. Based on the reported equilibrium dialysis data and in conjunction with other experimental data, a model for the binding of Ca(II) and Mg(II) to bovine prothrombin fragment 1 is proposed. The key difference between the binding of these divalent ions is that Ca(II) apparently causes a specific conformational change reflected by the cooperativity observed in the Ca(II) titration. The binding of Ca(II) ions to the three tight, cooperative sites establishes a conformation that is essential for phospholipid X Ca(II) X protein binding. The filling of the loose sites with Ca(II) ions leads to charge reduction and subsequent phospholipid X Ca(II) X protein complex interaction. Binding of Mg(II) to bovine prothrombin fragment 1 does not yield a complex with the necessary phospholipid-binding conformation. However, Mg(II) is apparently capable of stabilizing the Ca(II) conformation as is observed in the mixed metal ion binding data and the synergism in thrombin formation.  相似文献   

2.
M E Lee  T Nowak 《Biochemistry》1992,31(7):2172-2180
A new, more gentle enzyme purification for yeast enolase was developed. A series of kinetic experiments was performed with yeast enolase where the concentration of Mg(II) is kept constant and at the Km' level; the addition of Mn(II), Zn(II), or Cu(II) gives a hyperbolic decrease in the enzyme activity. The final velocity of these mixed-metal systems is the same as the velocity obtained only with Mn(II), Zn(II), or Cu(II), respectively. The concentration of the second metal that gives half-maximal effect in the presence of Mg(II) is approximately the same as the apparent Km (Km') value measured for that cation alone. Direct binding of Mn(II) to apoenolase in the absence and presence of Mg(II) shows that Mn(II) and Mg(II) compete for the same metal site on enolase. In the presence of D-2-phosphoglycerate (PGA) and Mg(II), only a single cation site per monomer is occupied by Mn(II). Water proton relaxation rate (PRR) studies of enzyme-ligand complexes containing Mn(II) and Mn(II) in the presence of Mg(II) are consistent with Mn(II) binding at site I under both conditions. PRR titrations of ligands such as the substrate PGA or the inhibitors orthophosphate or fluoride to the enolase-Mn(II)-Mg(II) complex are similar to those obtained for the enolase-Mn(II) complex, also indicating that Mn(II) is at site I in the presence of Mg(II). High-resolution 1H and 31P NMR was used to determine the paramagnetic effect of enolase-bound Mn(II) on the relaxation rates of the nuclei of the competitive inhibitor phosphoglycolate. The distances between the bound Mn(II) and the nuclei were calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Protein of regenerating liver (PRL)-1, -2, and -3 comprise a subgroup of closely related protein-tyrosine phosphatases featuring a C-terminal prenylation motif conforming to either the consensus sequence for farnesylation, CAAX, or geranylgeranylation, CCXX. Yeast two-hybrid screening for PRL-2-interacting proteins identified the beta-subunit of Rab geranylgeranyltransferase II (betaGGT II). The specific interaction of betaGGT II with PRL-2 but not with PRL-1 or -3 occurred in yeast and HeLa cells. Chimeric PRL-1/-2 molecules were tested for their interaction with betaGGT II, and revealed that the C-terminal region of PRL-2 is required for interaction, possibly the PRL variable region immediately preceeding the CAAX box. Additionally, PRL-2 prenylation is prequisite for betaGGT II binding. As prenylated PRL-2 is localized to the early endosome, we propose that this is where the interaction occurs. PRL-2 is not a substrate for betaGGT II, as isoprenoid analysis showed that PRL-2 was solely farnesylated in vivo. Co-expression of the alpha-subunit (alpha) of GGT II, betaGGT II, and PRL-2 resulted in alpha/betaGGT II heterodimer formation and prevented PRL-2 binding. Expression of PRL-2 alone inhibited the endogenous alpha/betaGGT II activity in HeLa cells. Together, these results indicate that the binding of alphaGGT II and PRL-2 to betaGGT II is mutually exclusive, and suggest that PRL-2 may function as a regulator of GGT II activity.  相似文献   

4.
The decatenation activity of DNA topoisomerase II is essential for viability as eukaryotic cells traverse mitosis. Phosphorylation has been shown to stimulate topoisomerase II activity in vitro. Here we show that topoisomerase II is a phosphoprotein in yeast and that the level of incorporated phosphate is significantly higher at mitosis than in G1. Comparison of tryptic phosphopeptide maps reveals that the major phosphorylation sites in vivo are targets for casein kinase II. Incorporation of phosphate into topoisomerase II is nearly undetectable at the non-permissive temperature in a conditional casein kinase II mutant. The sites modified by casein kinase II are located in the extreme C-terminal domain of topoisomerase II. This domain is absent in prokaryotic and highly divergent among eukaryotic type II topoisomerases, and may serve to regulate functions of topoisomerase II that are unique to eukaryotic cells.  相似文献   

5.
Enzyme elements that are involved in the reversible cyclization of L-carbamylaspartate to L-dihdroorotate catalyzed by dihydroorotase (EC 3.5.2.3) from Clostridium oroticum (ATCC 25750) have been studied. Removal of Zn(II) from the enzyme by chelators followed by incubation of apoenzyme with Co(II) results in replacement of two to three of the four Zn(II) ions per molecule by Co(II). The catalytic properties of the Zn(II)Co(II) dihydroorotase are different from those of native enzyme. The Vmax is increased for both the synthesis and hydrolysis of L-dihydroorotate. The Km for L-dihydroorotate is unchanged, while the Km for L-carbamylaspartate is increased more than twofold. On the other hand, the kinetic properties of Zn(II)-reconstituted dihydroorotase are indistinguishable from those of native enzyme. The pH dependence of Vmax is also altered by the Co(II) substitution. For both Zn(II)- and Zn(II)Co(II)-dihydroorotase, this pH dependence is well described by a single ionization and the pK's for L-dihydroorotate synthesis and hydrolysis are different. Substitution with Co(II) increases the pK for both reaction directions to different extents. These results strongly support a role for the tightly bound metals in the catalytic mechanism. In addition, diethylpyrocarbonate rapidly inactivates the enzyme. The inactivation is prevented by L-dihydroorotate. This result is consistent with a role for at least one histidine in catalysis. The possibility that C. oroticum dihydroorotase may be useful model for the more complex mammalian enzyme is considered.  相似文献   

6.
The ATP-independent type I and the ATP-dependent type II DNA topoisomerase of the yeast Saccharomyces cerevisiae have been purified to near homogeneity, and the purification procedures are reported. Both purified topoisomerases are single subunit enzymes with monomer weights of Mr = 90,000 and 150,000 for the type I and type II enzyme, respectively. Sedimentation and gel filtration data suggest that the type I enzyme is monomeric and the type II enzyme is dimeric. Similar to other purified eukaryotic topoisomerases, the yeast type I enzyme does not require a divalent cation for activity, but is stimulated 10-20-fold in the presence of 7-10 mM Mg(II) or Ca(II). Mn(II) is about 25% as efficient as Mg(II) in this stimulation but Co(II) is inhibitory. The yeast type II topoisomerase has an absolute requirement for a divalent cation: Mg(II) is the most effective, whereas Mn(II), Ca(II), or Co(II) supports the reaction to a lesser extent. The type II enzyme also requires ATP or dATP; the nonhydrolyzable ATP analogues adenylyl imidodiphosphate and adenylyl (beta,gamma-methylene)diphosphonate are potent inhibitors. Both yeast topoisomerases are completely inhibited by N-ethylmaleimide at 0.5 mM. In addition, the type II enzyme, but not the type I enzyme, is inhibited to various extents by coumermycin, ethidium, and berenil. Both topoisomerases are nuclear enzymes; no topoisomerase specific to mitochondria has been detected.  相似文献   

7.
A family of aspartic proteases called plasmepsins is important for hemoglobin degradation in intraerythrocytic Plasmodium parasites. Plasmepsin II (PM II) is the best studied member of this family. PM II and its close orthologs and paralogs form homodimers with extensive interfaces in all known crystal structures. This raised the question whether the homodimer is the functional subunit of plasmepsins in solution. We have used gel filtration chromatography, site-directed mutagenesis, and analytical ultracentrifugation to study the oligomeric status of PM II in solution. Our results reveal that PM II exists mainly as a monomer in solution and that the monomer is fully functional for catalysis. A hydrophobic loop at the PM II monomer surface, which would be buried in a PM II dimer, is shown to be essential for the hemoglobin degradation capability of PM II.  相似文献   

8.
The effect of Cu(II), Ni(II), Zn(II), Mg(II), and Mn(II) on the fluorescence of porcine kidney cytosol leucine aminopeptidase and three of its dansyl(Dns) peptide substrates, Leu-Gly-NHNH-Dns, Leu-Gly-NH(CH2)2NH-Dns, and Leu-Gly-NH(CH2)6NH-Dns, has been investigated. These five metal ions were chosen for study because each binds to the regulatory metal binding site of leucine aminopeptidase. Since the binding is relatively weak, kinetic studies of the different metalloderivatives of the enzyme are normally carried out in the presence of large molar excesses of these metal ions that can potentially affect both the enzyme and substrate. The fluorescence of all of the dansyl-peptides, as well as several other dansyl species, is quenched by Ni(II) and Cu(II), but not by Mg(II), Mn(II), or Zn(II). The absorption spectra of these dansyl substrates are also perturbed by Ni(II) and Cu(II). The rate at which maximal quenching for some dansyl species is attained after mixing with Ni(II) and Cu(II) is slow and the quenching is reversed on addition of EDTA. These results indicate that the quenching is the result of complex formation between the fluorophores and these metal ions. The association constants for the metal complexes have been determined from Stern-Volmer plots. In addition to complex formation, Ni(II) and Cu(II) cause the degradation of Leu-Gly-NHNH-Dns through a two step mechanism involving loss of dansic acid. Ni(II) and Cu(II) also partially quench the fluorescence of leucine aminopeptidase through contact with its surface accessible Trp residues. These observations indicate that care must be taken in stopped flow fluorescence studies of reactions between this enzyme and its dansyl substrates to avoid adverse effects brought about by Ni(II) and Cu(II).  相似文献   

9.
The kinetics and mechanism for Ni(II)-transfer of the native sequence tripeptide, L-aspartyl-L-alanyl-L-histidine-N-methylamide (AAHNMA), representing the Ni(II)-transport site of human serum albumin (HSA) and L-histidine (L-His) was studied in forward and reverse reactions in the pH range 6.5 to 9.0 at I = 0.2 and 25 degrees C. For the Ni(II)-transfer from Ni(II)-(L-His)2 to native sequence peptide, the rate-determining step is the formation of a mixed-ligand complex of NiH-1AB by deprotonation of peptide nitrogen from NiAB where A and B denote the anionic forms of AAHNMA and L-His, respectively. For the Ni(II)-transfer from Ni(II)-peptide to L-His, the rate-determining step is a bond breaking between Ni(II) and peptide nitrogen to form NiH-1A by protonation to a peptide nitrogen of NiH-2A. The equilibrium constants for the metal-transfer reaction of MH-2A + 2HB in equilibrium MB2 + A (A = Ni(II), Cu(II] were 10(3.29) and 10(0.78) for Ni(II) and Cu(II), respectively. NiB2 is 324 times as stable as CuB2. Furthermore, the ratio of Ni(II)/Cu(II) in the rate constants for the reaction of MB2 with A was found to be 2.8 x 10(-4). Thus, despite the similarities of Cu(II) and Ni(II) in the metal-binding sites of HSA and in reaction mechanism, Ni(II)-(L-His)2 complex is so stable thermodynamically and kinetically, compared to the Cu(II)-(L-His)2 complex, that Ni(II) is hardly transferred from Ni(II)-(L-His)2 to native sequence peptide. These findings may support specificities in the Ni(II)-transfer, its organ distribution, and its excretion through urine in vivo.  相似文献   

10.
11.
Dendritic cells (DCs) express major histocompatibility complex class II (MHC II) to present peptide antigens to T cells. In immature DCs, which bear low cell surface levels of MHC II, peptide-loaded MHC II is ubiquitinated. Ubiquitination drives the endocytosis and sorting of MHC II to the luminal vesicles of multivesicular bodies (MVBs) for lysosomal degradation. Ubiquitination of MHC II is abrogated in activated DCs, resulting in an increased cell surface expression. We here provide evidence for an alternative MVB sorting mechanism for MHC II in antigen-loaded DCs, which is triggered by cognately interacting antigen-specific CD4+ T cells. At these conditions, DCs generate MVBs with MHC II and CD9 carrying luminal vesicles that are secreted as exosomes and transferred to the interacting T cells. Sorting of MHC II into exosomes was, in contrast to lysosomal targeting, independent of MHC II ubiquitination but rather correlated with its incorporation into CD9 containing detergent-resistant membranes. Together, these data indicate two distinct MVB pathways: one for lysosomal targeting and the other for exosome secretion.  相似文献   

12.
We have previously shown that carbonic anhydrase II usually described as a cytoplasmic resident isoform (cCAH II) is secreted by the rat coagulating gland (sCAH II) via the apocrine secretion mode. To get more detailed information why CAH II is cytoplasmic resident in some organs and secreted in others we cloned and sequenced the cDNA of rat coagulating gland sCAH II. The sequence of the secretory form was found to be completely identical with the cCAH II. Therefore, a signal peptide targeting sCAH II for apocrine secretion can be excluded. Considering the fact that other apocrine secreted proteins are glycosylated, cCAH II and sCAH II were analyzed for carbohydrate substitutions. As expected for a cytoplasmic protein, no glycan modification could be identified in cCAH II. In contrast, sCAH II carried exclusively Gal, GlcNAc and Fuc residues in a molar ratio of 1:0.8:0.5. Carbohydrate linkage analyses demonstrated the presence of terminal Fuc, terminal, 3-substituted and 3,6-disubstituted Gal as well as 4-substituted and 3,4-disubstituted GlcNAc. The composition of the glycan constituents as well as deglycosylation experiments clearly proved that sCAH II carries neither conventional mammalian-type N-glycans nor mucin-type O-linked sugar chains. Lacking a signal peptide for ER translocation, glycosylation of sCAH II must occur within the cytoplasmic compartment. Further studies have to elucidate whether or not glycosylation of sCAH II is essential for the apocrine release of the protein.  相似文献   

13.
14.
15.
Gerald T. Babcock  Kenneth Sauer 《BBA》1973,325(3):483-503
An analysis of electron paramagnetic resonance Signal II in spinach chloroplasts has been made using both continuous and flashing light techniques. In order to perform the experiments we developed a method which allows us to obtain fresh, untreated chloroplasts with low dark levels of Signal II. Under these conditions a single 10-μs flash is sufficient to generate greater than 80% of the possible light-induced increase in Signal II spin concentration. The risetime for this flash-induced increase in Signal II is approx. 1 s. The close association of Signal II with Photo-system II is confirmed by the observations that red light is more effective than is far red light in generating Signal II, and that 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) does not inhibit the formation of the radical. Single flash saturation curves for the flash-induced increase in Signal I and Signal II indicate that the quantum efficiency for Signal II formation is close to that for Signal I. While one or two flashes (spaced 10 ms apart) are quite efficient in generating Signal II, three or four flashes are much less effective. However, if this spacing is decreased to 100 μs, three or four flashes become as efficient as one or two flashes. From observations of a deficiency of O2 evolved during the initial flashes of dark-adapted chloroplasts, we conclude that the species which gives rise to Signal II is able to compete with water for oxidizing equivalents generated by Photosystem II. On the basis of these results we postulate a model in which Signal II arises from an oxidized radical which is produced by a slow electron transfer to the specific states S2 and S3 on the water side of Photo-system II.  相似文献   

16.
The Zn(II) site of the dimeric Cu(II),Zn(II)-superoxide dismutase from Saccharomyces cerevisiae has been examined by means of perturbed angular correlation of gamma rays (PAC) on the Cu(II),Cd(II)- and Cu(I),Cd(II)-superoxide dismutase. The PAC spectrum for the Cu(II),Cd(II) enzyme reveals two different, pH independent, coordination geometries for the Cd site. Removal of Cu(II) does not affect the PAC spectrum, which suggests that Cu(II) and Cd(II) do not share a common histidine side chain as ligand. The results are consistent with either an equilibrium between two coordination geometries for Cd(II) in each subunit or a difference in the structure of the Cd(II) site in the two subunits. In contrast, in the reduced enzyme only one structure is present, identical for the two subunits.  相似文献   

17.
SlyD is a Ni(II)-binding protein that contributes to nickel homeostasis in Escherichia coli. The C-terminal domain of SlyD contains a rich variety of metal-binding amino acids, suggesting broader metal binding capabilities, and previous work demonstrated that the protein can coordinate several types of first-row transition metals. However, the binding of SlyD to metals other than Ni(II) has not been previously characterized. To improve our understanding of the in vitro metal-binding activity of SlyD and how it correlates with the in vivo function of this protein, the interactions between SlyD and the series of biologically relevant transition metals [Mn(II), Fe(II), Co(II), Cu(I), and Zn(II)] were examined by using a combination of optical spectroscopy and mass spectrometry. Binding of SlyD to Mn(II) or Fe(II) ions was not detected, but the protein coordinates multiple ions of Co(II), Zn(II), and Cu(I) with appreciable affinity (K(D) values in or below the nanomolar range), highlighting the promiscuous nature of this protein. The order of affinities of SlyD for the metals examined is as follows: Mn(II) and Fe(II) < Co(II) < Ni(II) ~ Zn(II) ? Cu(I). Although the purified protein is unable to overcome the large thermodynamic preference for Cu(I) and exclude Zn(II) chelation in the presence of Ni(II), in vivo studies reveal a Ni(II)-specific function for the protein. Furthermore, these latter experiments support a specific role for SlyD as a [NiFe]-hydrogenase enzyme maturation factor. The implications of the divergence between the metal selectivity of SlyD in vitro and the specific activity in vivo are discussed.  相似文献   

18.
High affinity binding sites for angiotensin II in bovine and rat brain membranes have been identified and characterized using monoiodinated Ile5-angiotensin II of high specific radioactivity. Degradation of labeled and unlabeled peptide by washed brain particulate fractions was prevented by adding glucagon to the final incubation medium and including a proteolytic enzyme inhibitor (phenylmethylsulfonyl fluoride) in preincubation and incubation procedures. 125I-Angiotensin II binding can be studied using either centrifugation or filtration techniques to separate tissue-bound radioactivity. 125I-Angiotensin II binding to calf brain membranes is saturable and reversible, with a dissociation binding constant of 0.2 nM at 37 degrees. A similar binding constant is found in rat brain membranes. Analogues and fragments of angiotensin II compete for these brain binding sites with potencies which correlate with both their in vivo potencies and their binding inhibition protencies at adrenal cortex angiotensin II receptors. Angiotensin I is 1 to 2 orders of magnitude weaker than angiotensin II; the 3-8 hexapeptide and 4-8 pentapeptide are much weaker still. (desAsp1) angiotensin II (angiotensin III) is slightly more potent than angiotensin II, as are several antagonists of angiotensin II with aliphatic amino acids substituted at position 8. In calf brain 125I-angiotensin II binding is restricted almost exclusively to the cerebellum (cortex and deep nuclei). In rat brain, angiotensin II binding is highest in the thalamus-hypothalamus, midbrain, and brainstem, areas which are believed to be involved in mediating angiotensin II-induced central effects. These findings illustrate the presence of high affinity specific binding sites for angiotensin II in rat and bovine brain and suggest a physiological role for angiotensin peptides in the central nervous system.  相似文献   

19.
Fundamental issues in zinc biology are how proteins control the concentrations of free Zn(II) ions and how tightly they interact with them. Since, basically, the Zn(II) stability constants of only two cytosolic zinc enzymes, carbonic anhydrase and superoxide dismutase, have been reported, the affinity for Zn(II) of another zinc enzyme, sorbitol dehydrogenase (SDH), was determined. Its log K is 11.2 +/- 0.1, which is similar to the log K values of carbonic anhydrase and superoxide dismutase despite considerable differences in the coordination environments of Zn(II) in these enzymes. Protein tyrosine phosphatase 1B (PTP 1B), on the other hand, is not classified as a zinc enzyme but is strongly inhibited by Zn(II), with log K = 7.8 +/- 0.1. In order to test whether or not metallothionein (MT) can serve as a source for Zn(II) ions, it was used to control free Zn(II) ion concentrations. MT makes Zn(II) available for both PTP 1B and the apoform of SDH. However, whether or not Zn(II) ions are indeed available for interaction with these enzymes depends on the thionein (T) to MT ratio and the redox poise. At ratios [T/(MT + T) = 0.08-0.31] prevailing in tissues and cells, picomolar concentrations of free Zn(II) are available from MT for reconstituting apoenzymes with Zn(II). Under conditions of decreased ratios, nanomolar concentrations of free Zn(II) become available and affect enzymes that are not zinc metalloenzymes. The match between the Zn(II) buffering capacity of MT and the Zn(II) affinity of proteins suggests a function of MT in controlling cellular Zn(II) availability.  相似文献   

20.
Transition row metal ions are both essential and toxic to microorganisms. Zinc in excess has significant toxicity to bacteria, and host release of Zn(II) at mucosal surfaces is an important innate defence mechanism. However, the molecular mechanisms by which Zn(II) affords protection have not been defined. We show that in Streptococcus pneumoniae extracellular Zn(II) inhibits the acquisition of the essential metal Mn(II) by competing for binding to the solute binding protein PsaA. We show that, although Mn(II) is the high-affinity substrate for PsaA, Zn(II) can still bind, albeit with a difference in affinity of nearly two orders of magnitude. Despite the difference in metal ion affinities, high-resolution structures of PsaA in complex with Mn(II) or Zn(II) showed almost no difference. However, Zn(II)-PsaA is significantly more thermally stable than Mn(II)-PsaA, suggesting that Zn(II) binding may be irreversible. In vitro growth analyses show that extracellular Zn(II) is able to inhibit Mn(II) intracellular accumulation with little effect on intracellular Zn(II). The phenotype of S. pneumoniae grown at high Zn(II):Mn(II) ratios, i.e. induced Mn(II) starvation, closely mimicked a ΔpsaA mutant, which is unable to accumulate Mn(II). S. pneumoniae infection in vivo elicits massive elevation of the Zn(II):Mn(II) ratio and, in vitro, these Zn(II):Mn(II) ratios inhibited growth due to Mn(II) starvation, resulting in heightened sensitivity to oxidative stress and polymorphonuclear leucocyte killing. These results demonstrate that microbial susceptibility to Zn(II) toxicity is mediated by extracellular cation competition and that this can be harnessed by the innate immune response.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号